Next Article in Journal
Improving the Corrosion Behavior of Biodegradable AM60 Alloy through Plasma Electrolytic Oxidation
Next Article in Special Issue
Dual Cluster Model for Medium-Range Order in Metallic Glasses
Previous Article in Journal
Mechanism of Electromagnetic Field in the Sub-RapidSolidification of High-Strength Al-Cu-Li Alloy Produced by Twin-Roll Casting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Co/Ni Substituting Fe on Magnetocaloric Properties of Fe-Based Bulk Metallic Glasses

1
Xinjiang Key Laboratory of Solid State Physics and Devices, Xinjiang University, Urumqi 830046, China
2
School of Physics Science and Technology, Xinjiang University, Urumqi 830046, China
3
School of Materials Science and Engineering, Tianjin University, Tianjin 300072, China
4
Key Laboratory of Magnetic Materials and Devices, Ningbo Institute of Materials Technology & Engineering, Chinese Academy of Sciences, Ningbo 315201, China
5
School of Mechanical Engineering, Dongguan University of Technology, Dongguan 523808, China
6
State Key Laboratory for Advanced Metals and Materials, University of Science and Technology Beijing, Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Metals 2021, 11(6), 950; https://doi.org/10.3390/met11060950
Submission received: 15 May 2021 / Revised: 3 June 2021 / Accepted: 7 June 2021 / Published: 11 June 2021
(This article belongs to the Special Issue Structure and Properties of Amorphous Metallic Alloys)

Abstract

:
In this work, Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) bulk metallic glasses (BMGs) were prepared, and the effect of the Co/Ni elements substitution for Fe on the magnetocaloric properties of Fe80P13C7 BMG has been investigated systematically. The Curie temperature (TC) of the present Fe-based BMGs increases with the substitution of Fe by Co/Ni. The magnetic entropy change ( Δ S M ) of the present Fe-based BMGs increases first and then decreases with the increase of Fe substituted by Co, but monotonically decreases with the increase of Fe substituted by Ni. Among the present Fe-based BMGs, the Fe75Co5P13C7 BMG exhibits the maximum Δ S M value of 5.21 J kg−1 K−1 at an applied field of 5 T, which is the largest value among Fe-based amorphous alloys without any rare earth elements reported so far. The present Fe-based BMGs exhibit the large glass forming ability, tunable TC and enhanced Δ S M value, which are beneficial for magnetic refrigerant materials.

1. Introduction

Compared with the traditional refrigeration technology of compressed gas medium, the magnetic refrigeration technology [1,2] is an environmentally friendly technology with high efficiency, low energy consumption and no environmental pollution [3,4]. Up to now, a series of the crystalline compounds, for example, Gd-Si-Ge [5,6], La-Fe-Si [7] and Mn-Fe-P-Ps [8], etc., show a giant magnetic entropy change ( Δ S M ) based on the first order magnetic phase transition (FOMT), but also some shortcomings including the low corrosion resistance, high hysteresis losses and expensive raw materials, which greatly limit the application of these materials in the field of magnetic refrigeration [9,10]. Compared with the FOMT magnetocaloric materials, metallic glasses, e.g., Fe-based amorphous alloys, exhibit magnetocaloric effects based on the second order magnetic phase transition (SOMT) and possess the unique advantages of magnetic refrigeration materials, such as superior mechanical properties, low hysteresis and thermal hysteresis, higher electrical resistivity, and so on [11,12]. Moreover, Fe-based amorphous alloys show great potential for commercial use because of their abundant raw materials. However, Fe-based amorphous alloys show a low value of Δ S M frequently, which greatly restricts their application as magnetic refrigeration materials [13]. Therefore, it is of significance and importance for Fe-based amorphous alloys to improve magnetocaloric properties.
Many studies have indicated that the element addition to Fe-based amorphous alloys may cause significant changes in their magnetocaloric properties. Usually, the rare earth elements are chosen to be a regulating element and are added into Fe-based amorphous alloys to improve magnetocaloric properties due to their large atomic magnetic moments. Recent studies show that minor additions of Gd or Dy elements increase the value of Δ S M while the excessive addition of Gd or Dy causes the value of Δ S M to decrease [14]. However, Ce element addition leads to the decrease of Δ S M [15] in Fe-Si-Nb-B-Cu metallic glasses. The addition of transition metals usually increases the TC and Δ S M of Fe-based amorphous alloys due to the stronger atomic exchange interaction. The addition of Cu improves the magnetocaloric properties of Fe-Zr amorphous alloys [16]. The partial substitution of Fe by Co and Ni in Fe-Zr-B-Cu amorphous alloy leads to the increase of TC, and appropriate doping Co and Ni elements enhances the value of Δ S M for Fe71.5Co8.25Ni8.25Zr7B4Cu1 amorphous alloy [17]. These results indicate that systematic investigation of the effect of transition metals on magnetocaloric effect should be significant.
Previous studies showed that Fe80P13C7 bulk metallic glass (BMG) exhibited excellent magnetocaloric performance with the Δ S M of 5.05 J kg−1 K−1 under an applied field of 5 T [18]. To further increase the Δ S M value of Fe80P13C7 BMG, the glass forming ability (GFA), magnetic properties and magnetocaloric performance of Fe80P13C7 BMG affected by partially substituting the transition metals (Co/Ni) for Fe were investigated systematically.

2. Experiment

Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) alloy ingots were prepared by torchmelting the high purity elements Fe powders (99.9% pure), Fe3P Powders (99.5% pure), graphite powders (99.9% pure), cobalt powders (99.9% pure) and nickel powders (99.7% pure) with a torch in a clean quartz tube under a high-purity argon atmosphere. The ingots were then placed into clean quartz tubes containing a fluxing agent of B2O3 and CaO powder (mass ratio = 3:1) and fluxed in a vacuum at 1150 degrees for 4 h. After that, the glass alloy rods with a maximum diameter of 2.5 mm and a length of several centimeters were produced by the J-quenching technique [18].
The structure of the alloy rods was measured by Burker D8 advance X-ray diffractometer (XRD) (Bruker, Billerica, MA, USA) with Cu Kα radiation. The rod samples were smashed to small pieces by a hammer before the XRD test to obtain the pattern with good quality. The thermal behavior of the glassy rod was estimated by NETZSCH DSC 404F1 differential scanning calorimetry (DSC) (NETZSCH, Selb, Germany) with a heating rate of 0.33 K/s under an Ar atmosphere. The magnetic hysteresis loop of glassy rod samples was tested with the maximum applied field up to 10,000 Oe at the room temperature by a Lake Shore 7400 vibrating sample magnetometer (VSM) (Lake Shore, Carson, CA, USA). The field and temperature dependences of magnetization curves of the glassy rods were tested by Quantum Design® superconducting quantum interference device magnetometer (MPMS) (Quantum Design Int., San Diego, CA, USA), in which the measurement specimens were first cut about 3 mm from the bulk glassy rods with a diameter of 1 mm and then ground into rectangular slices with around 40 µm thick and 1 mm width. The temperature interval is 10 K away from TC and 5 K near the TC in the isothermal magnetization curve measurements. The density of the samples was determined by Archimedes’ method. Except for XRD test, the samples used in other tests are all 1 mm in diameter to avoid the risk of partial crystallization.

3. Result and Discussion

The XRD patterns of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) alloy rods with the corresponding maximum diameters (Dmax) for fully glass formation [19,20] are shown in Figure 1. There are only typical halo peaks and no crystalline peaks in the XRD patterns of all the samples, verifying the fully glassy phase formation. Figure 2 shows the DSC curves of the as-cast glassy alloy rod specimen. All the curves show the sequential transition of the glass transition, supercooled liquid region and two-step crystallization. The DSC curves can further confirm the glassy structure of the present samples. With the substitution content of both Co and Ni for Fe increases from 0 to 10 at.%, the Dmax of the present Fe-based BMGs increases from 2.0 to 2.5 mm, indicating that substituting Fe with a minor amount of Co/Ni can increase the GFA of Fe80P13C7 alloy to a certain extent.
Figure 3 shows the hysteresis loops of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy rods tested by VSM, and the determined saturation magnetization (Js) of the measured samples are shown in Table 1. All the glassy rod samples show good soft magnetic performance and low coercivity. The Js of the glassy rod samples firstly increases from 160 to 165 emu/g with the increase of Co substituting for Fe from 0 to 5 at.%, and then decreases to 157 emu/g when the Co content reaches 10 at.%, while the Js value monotonically decreases from 160 to 150 emu/g with the Ni substitution for Fe increases from 0 to 10 at.%.
Figure 4a shows that the temperature varies with magnetization (M-T curve) of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy rod samples tested by MPMS at a magnetic applied field of 0.02 T, and Figure 4b shows the TC determined by the differentiation of M-T curves and summarized in Table 1. It can be concluded that with the increase of Co/Ni substituting Fe, the TC of Fe-based BMGs increases, and further the rise rate of TC with the Co content is greater than that with the Ni content. Based on the mean field theory, TC is proportionate to the saturation magnetization and the exchange integral, in which the exchange integral is dependent on the distance between magnetic atoms [21]. The radii of Fe, Co and Ni atoms are 0.124 nm, 0.125 nm and 0.125 nm, respectively [22]. Therefore, based on the Bathe-Slater curve [21], the replacement of Co/Ni for part of Fe caused a stronger atomic exchange interaction, and thus the higher TC.
The isothermal magnetization curves of the Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rod samples at the magnetic applied fields of 1.5 to 5 T are shown in Figure 5. When the test temperature is below TC, the samples reach the magnetization saturation under low applied magnetic fields, but when the temperature gets close to or above TC, the M-H curves of the samples become linear, indicating that the samples turn from ferromagnetic to paramagnetic state [23].
In most research on characterization of magnetocaloric materials, the magnetic entropy change ( Δ S M ) is an important parameter and Δ S M under the applied maximum magnetic field of H m a x is defined as the follow equation [3,24]:
Δ S M = 0 H m a x ( M T ) H d H
where M is the magnetization under the applied magnetic field of H. Equation (1) can also be written in other terms:
Δ S M   ( T i , H m a x ) = 0 H m a x M ( T i , H ) d H 0 H m a x M ( T i + 1 ,     H ) d H T i T i + 1
where M ( T i , H ) and M ( T i + 1 , H ) are the magnetization at the temperature of T i and T i + 1 , respectively, and under the applied magnetic field of H. Based on the M-H curves of the samples at different temperature in Figure 6, the change of − Δ S M with temperature of Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy rod samples under the different applied maximum fields can be constructed according to Formula (2) and is shown in Figure 6, and the peak values ( | Δ S M peak | ) determined from the Δ S M T curve of the samples under the applied field of 1.5 T and 5.0 T are summarized in Table 1. The | Δ S M peak | values of Fe80P13C7, Fe75Co5P13C7, Fe70Co10P13C7, Fe75Ni5P13C7 and Fe70Ni10P13C7 are 2.20, 2.34, 2.08, 2.04 and 1.46 J kg−1 K−1 at an applied field of 1.5 T, respectively, and 5.05, 5.21, 4.98, 4.60 and 3.26 J kg−1 K−1 at an applied field of 5.0 T, respectively. Compared to the Fe80P13C7 BMG, replacing 5 at.% of Fe by Co increases the value of | Δ S M peak | while 10 at.% substitution of Co for Fe reduces the value of | Δ S M peak | ; the value of | Δ S M peak | continuously decreases with the increasing of Ni substituted for Fe content. It can be found that the changing trend of | Δ S M peak | value is consistent with that of Js value with the substitution of Co/Ni for Fe in the present Fe-based BMGs. Many studies [24] have pointed out that there is a positive correlation between the Js of magnetic materials and the | Δ S M peak | . The | Δ S M peak | as a function of Js for the present Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) BMGs at the magnetic applied field of 1.5 T and 5.0 T are shown in Figure 7, in which the good linear relationship with the R-square value of 0.864, indicates that the present Fe-based BMGs follow this correlation relationship.
Another important parameter that is often applied for the characterization of magnetocaloric effect is refrigeration capacity (RC), which equals to the area of the Δ S M T curve. The R C F W H M can be obtained by the following equation [17,25]:
R C F W H M = Δ S M peak × Δ T F W H M
where Δ T F W H M is the temperature interval corresponding to the half-width of the Δ S M  − T curve. The Δ T F W H M and R C F W H M values of each glassy rod samples are listed in Table 1. We can see that the R C F W H M of the Fe80P13C7, Fe75Co5P13C7, Fe70Co10P13C7, Fe75Ni5P13C7, Fe70Ni10P13C7 glassy rods are 125.6, 91.6, 115.2, 106.5, 71.1 J kg1 under 1.5 T, respectively. It is worth noting that, compared with the Fe80P13C7 BMG, the Fe80-xCoxP13C7 BMG shows the larger | Δ S M peak | but smaller R C F W H M , which is attributed to the smaller Δ T F W H M of the Fe75Co5P13C7 BMG, as shown in Table 1. The smaller Δ T F W H M is presumably the stronger magnetic coupling in the Fe75Co5P13C7 BMG.

4. Conclusions

In this work, the Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) bulk glassy rods are fabricated by combining fluxing treatment and the J-quenching technique. Moreover, their magnetocaloric properties are investigated. The replacement of Co/Ni for Fe caused the increase of TC from 579 K for x = 0 to 649 K for M = Co, x = 10 and 600 K for M = Ni, x = 10. The | Δ S M peak | and R C F W H M values of Fe80P13C7, Fe75Co5P13C7, Fe70Co10P13C7, Fe75Ni5P13C7, Fe70Ni10P13C7 are 5.05, 5.21, 4.98, 4.60, 3.26 J kg−1 K−1, and 479.8, 331.6, 462.1, 381.8, 211.2 J kg−1, respectively, at the magnetic applied field of 5 T. Significantly, the Fe75Co5P13C7 amorphous alloy exhibits the largest value of | Δ S M peak | in the Fe-based amorphous alloys without any rare earth elements as reported so far. The high TC of the present BMGs may limit their application but can be lowered to room temperature by adding some selected elements, which makes them possible for room temperature magnetic refrigeration.

Author Contributions

Conceptualization, J.G. and Q.L.; formal analysis, J.G.; investigation, J.G. and L.X.; resources, L.X., Q.L., J.H. and X.M.; writing—original draft preparation, J.G.; writing—review and editing, C.L., Q.L., J.H., C.C., H.L. and X.M.; funding acquisition, Q.L. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 51771161; the Leading Talents of Tianshan Cedar Program of Xinjiang Uygur Autonomous Region, grant number 2019XS02; the Tianshan Innovation Team Program of Xinjiang Uygur Autonomous Region, grant number 2020D14038; and the financial support from Beijing Municipal Natural Science Foundation, grant number 2202033.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This research was supported by National Natural Science Foundation of China (No. 51771161), the Leading Talents of Tianshan Cedar Program of Xinjiang Uygur Autonomous Region (No. 2019XS02), and the Tianshan Innovation Team Program of Xinjiang Uygur Autonomous Region (No. 2020D14038). The author (Hongxiang Li) appreciates the financial support from Beijing Municipal Natural Science Foundation (No. 2202033).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Franco, V.; Blázquez, J.S.; Ingale, A.; Conde, B. The magnetocaloric effect and magnetic refrigeration near room temperature: Materials and models. Annu. Rev. Mater. Res. 2012, 42, 305–342. [Google Scholar] [CrossRef] [Green Version]
  2. Gutfleisch, O.; Willard, M.A.; Brück, E.; Chen, C.H.; Sankar, S.G.; Liu, J.P. Magnetic materials and devices for the 21st century: Stronger, lighter, and more energy efficient. Adv. Mater. 2011, 23, 821–842. [Google Scholar] [CrossRef] [PubMed]
  3. Phan, M.-H.; Yu, S.-C. Review of the magnetocaloric effect in manganite materials. J. Magn. Magn. Mater. 2007, 308, 325–340. [Google Scholar] [CrossRef]
  4. Tishin, A.M.; Spichkin, Y.I. The Magnetocaloric Effect and Its Applications, 1st ed.; Institute of Physics: London, UK, 2003. [Google Scholar]
  5. Alves, C.S.; Gama, S.; Coelho, A.D.A.; Plaza, E.J.R.; Carvalho, A.M.G.; Cardoso, L.P.; Persiano, A.C. Giant magnetocaloric effect in Gd5 (Si2 Ge2) alloy with low purity Gd. Mater. Res. Mater. Res. 2004, 7, 535–538. [Google Scholar] [CrossRef]
  6. Kimel, A.V.; Kirilyuk, A.; Tsvetkov, A.; Pisarev, R.V.; Rasing, T. Laser-induced ultrafast spin reorientation in the antiferromagnet TmFeO3. Nature 2004, 429, 850–853. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, J.; Moore, J.D.; Skokov, K.P.; Krautz, M.; Löwe, K.; Barcza, A.; Katter, M.; Gutfleisch, O. Exploring La (Fe,Si)13-based magnetic refrigerants towards application. Scr. Mater. 2012, 67, 584–589. [Google Scholar] [CrossRef]
  8. Tegus, O.; Brück, E.; Buschow, K.H.J.; De Boer, F.R. Transition-metal-based magnetic refrigerants for room-temperature applications. Nat. Cell Biol. 2002, 415, 150–152. [Google Scholar] [CrossRef]
  9. Li, X.; Pan, Y. Magnetocaloric effect in Fe-Zr-B-M (M = Ni, Co, Al, and Ti) amorphous alloys. J. Appl. Phys. 2014, 116, 093910. [Google Scholar] [CrossRef]
  10. Wang, Y.F.; Qin, F.X.; Wang, Y.H.; Wang, H.; Das, R.; Phan, M.H.; Peng, H.X. Magnetocaloric effect of Gd-based microwires from binary to quaternary system. AIP Adv. 2017, 7, 056422. [Google Scholar] [CrossRef]
  11. Wang, W. The nature and properties of amorphous materials. Prog. Phys. 2013, 33, 177–350. [Google Scholar]
  12. Wang, F.; Inoue, A.; Han, Y. Excellent soft magnetic Fe-Co-B-based amorphous alloys with extremely high saturation magnetization above 1.85 T and low coercivity below 3 A/m. J. Alloys Compd. 2017, 711, 132–142. [Google Scholar] [CrossRef]
  13. Zhang, H.; Li, R.; Xu, T.; Liu, F.; Zhang, T. Near room-temperature magnetocaloric effect in FeMnPBC metallic glasses with tunable Curie temperature. J. Magn. Magn. Mater. 2013, 347, 131–135. [Google Scholar] [CrossRef]
  14. Zhao, L.; Tian, H.; Zhong, X.; Liu, Z.; Greneche, J.M.; Ramanujan, R.V. Influence of gadolinium and dysprosium substitution on magnetic properties and magnetocaloric effect of Fe78-xRExSi4Nb5B12Cu1 amorphous alloys. J. Rare Earths 2020, 38, 1317–1321. [Google Scholar] [CrossRef]
  15. Tian, H.; Zhong, X.; Liu, Z.; Zheng, Z.; Min, J.X. Achieving table-like magnetocaloric effect and large refrigerant capacity around room temperature in Fe78-xCexSi4Nb5B12Cu1 (x = 0−10) composite materials. Mater. Lett. 2015, 138, 64–66. [Google Scholar] [CrossRef]
  16. Yang, W.; Li, W.; Wan, C.; Huo, J.; Mo, J.; Liu, H.; Shen, B. Low-temperature magnetic properties and magnetocaloric effect of Fe-Zr-Cu amorphous alloys. J. Low Temp. Phys. 2020, 200, 51–61. [Google Scholar] [CrossRef]
  17. Caballero-Flores, R.; Franco, V.; Conde, A.; Knipling, K.E.; Willard, M.A. Influence of Co and Ni addition on the magnetocaloric effect in Fe88−2xCoxNixZr7B4Cu1 soft magnetic amorphous alloys. Appl. Phys. Lett. 2010, 96, 182506. [Google Scholar] [CrossRef]
  18. Yang, W.; Huo, J.; Liu, H.; Li, J.; Song, L.; Li, Q.; Xue, L.; Shen, B.; Inoue, A. Extraordinary magnetocaloric effect of Fe-based bulk glassy rods by combining fluxing treatment and J-quenching technique. J. Alloys Compd. 2016, 684, 29–33. [Google Scholar] [CrossRef]
  19. Xu, K.; Ling, H.; Li, Q.; Li, J.; Yao, K.; Guo, S. Effects of Co substitution for Fe on the glass forming ability and properties of Fe80P13C7 bulk metallic glasses. Intermetallics 2014, 51, 53–58. [Google Scholar] [CrossRef]
  20. Ma, X.H.; Zhang, L.; Yang, X.H.; Li, Q.; Huang, Y.D. Effect of Ni addition on corrosion resistance of FePC bulk glassy alloy. Corros. Eng. Sci. Technol. 2014, 50, 433–437. [Google Scholar] [CrossRef]
  21. Cullity, B.D.; Graham, C.D. Introduction to Magnetic Materials, 2nd ed.; John Wiley & Sons: Hoboken, NJ, USA, 2009. [Google Scholar]
  22. Takeuchi, A.; Inoue, A. Classification of bulk metallic glasses by atomic size difference, heat of mixing and period of constituent elements and its application to characterization of the main alloying element. Mater. Trans. 2005, 46, 2817–2829. [Google Scholar] [CrossRef] [Green Version]
  23. Kucuk, I.; Şarlar, K.; Adam, A.; Civan, E. Magnetocaloric and magnetoresistance properties in Co based (Co0.402Fe0.201Ni0.067B0.227Si0.053Nb0.05) 100-xCux (x = 0−1) glassy alloys. Philos. Mag. 2016, 96, 3120–3130. [Google Scholar] [CrossRef]
  24. Gschneidner Jr, K.A.; Pecharsky, V.K. Magnetocaloric materials. Annu. Rev. Mater. Sci. 2000, 30, 387–429. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, G.; Zhao, Z.; Zhang, X.; Li, H.; Zhao, G. Peculiar effect of rare earth doping on magnetic and magnetocaloric properties in Fe-rich amorphous ribbons. J. Alloys Compd. 2018, 735, 104–108. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) alloy rods with the corresponding maximum diameter for fully glass formation.
Figure 1. XRD patterns of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) alloy rods with the corresponding maximum diameter for fully glass formation.
Metals 11 00950 g001
Figure 2. DSC curves of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at a heating rate of 0.33 K/s.
Figure 2. DSC curves of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at a heating rate of 0.33 K/s.
Metals 11 00950 g002
Figure 3. The hysteresis loops of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at room temperature and the inset is the enlarged part of the magnetization curve.
Figure 3. The hysteresis loops of the as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at room temperature and the inset is the enlarged part of the magnetization curve.
Metals 11 00950 g003
Figure 4. Temperature varies with magnetization (a) and dM/dT (b) of as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rod samples.
Figure 4. Temperature varies with magnetization (a) and dM/dT (b) of as-cast Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rod samples.
Metals 11 00950 g004
Figure 5. The isothermal magnetization curves of Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at the magnetic applied fields up to 5 T.
Figure 5. The isothermal magnetization curves of Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at the magnetic applied fields up to 5 T.
Metals 11 00950 g005
Figure 6. Magnetic entropy changes as a function of temperature under 1.5–5 T for Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods.
Figure 6. Magnetic entropy changes as a function of temperature under 1.5–5 T for Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods.
Metals 11 00950 g006
Figure 7. The relationship between saturation magnetization (Js) at room temperature and the maximum magnetic entropy changes ( | Δ S M peak | ) of Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at a magnetic applied field of 1.5 T.
Figure 7. The relationship between saturation magnetization (Js) at room temperature and the maximum magnetic entropy changes ( | Δ S M peak | ) of Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) glassy alloy rods at a magnetic applied field of 1.5 T.
Metals 11 00950 g007
Table 1. The maximum diameter for fully glass formation (Dmax), magnetic transition temperature (TC), the saturation magnetization (Js) and magnetocaloric properties at the magnetic applied field of 1.5 T and 5 T for the present Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) BMGs.
Table 1. The maximum diameter for fully glass formation (Dmax), magnetic transition temperature (TC), the saturation magnetization (Js) and magnetocaloric properties at the magnetic applied field of 1.5 T and 5 T for the present Fe80-xMxP13C7 (M = Co, Ni; x = 0, 5 and 10 at.%) BMGs.
Composition
(at.%)
Dmax
(mm)
TC
(K)
Js
(emu/g)
−ΔSM
(J kg1 K1)
ΔTFWHM
(K)
RCFWHM
(J kg1)
1.5 T5 T1.5 T1.5 T5 T
Fe80P13C72.05811602.205.0557.1125.6479.8
Fe75Co5P13C72.36231652.345.2139.191.6331.6
Fe70Co10P13C72.56491572.084.9855.4115.2462.1
Fe75Ni5P13C72.55961552.044.6052.2106.5381.8
Fe70Ni10P13C72.56001501.463.2648.771.1211.2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, J.; Xie, L.; Liu, C.; Li, Q.; Huo, J.; Chang, C.; Li, H.; Ma, X. Effect of Co/Ni Substituting Fe on Magnetocaloric Properties of Fe-Based Bulk Metallic Glasses. Metals 2021, 11, 950. https://doi.org/10.3390/met11060950

AMA Style

Guo J, Xie L, Liu C, Li Q, Huo J, Chang C, Li H, Ma X. Effect of Co/Ni Substituting Fe on Magnetocaloric Properties of Fe-Based Bulk Metallic Glasses. Metals. 2021; 11(6):950. https://doi.org/10.3390/met11060950

Chicago/Turabian Style

Guo, Jia, Lei Xie, Cong Liu, Qiang Li, Juntao Huo, Chuntao Chang, Hongxiang Li, and Xu Ma. 2021. "Effect of Co/Ni Substituting Fe on Magnetocaloric Properties of Fe-Based Bulk Metallic Glasses" Metals 11, no. 6: 950. https://doi.org/10.3390/met11060950

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop